ArticlePDF Available

Coupling ATP utilization to protein remodeling by ClpB, a hexameric AAA+ protein

Authors:

Abstract and Figures

ClpB and Hsp104 are members of the AAA+ (ATPases associated with various cellular activities) family of proteins and are molecular machines involved in thermotolerance. They are hexameric proteins containing 12 ATP binding sites with two sites per protomer. ClpB and Hsp104 possess some innate protein remodeling activities; however, they require the collaboration of the DnaK/Hsp70 chaperone system to disaggregate and reactivate insoluble aggregated proteins. We investigated the mechanism by which ClpB couples ATP utilization to protein remodeling with and without the DnaK system. When wild-type ClpB, which is unable to remodel proteins alone in the presence of ATP, was mixed with a ClpB mutant that is unable to hydrolyze ATP, the heterohexamers surprisingly gained protein remodeling activity. Optimal protein remodeling by the heterohexamers in the absence of the DnaK system required approximately three active and three inactive protomers. In addition, the location of the active and inactive ATP binding sites in the hexamer was not important. The results suggest that in the absence of the DnaK system, ClpB acts by a probabilistic mechanism. However, when we measured protein disaggregation by ClpB heterohexamers in conjunction with the DnaK system, incorporation of a single inactive ClpB subunit blocked activity, supporting a sequential mechanism of ATP utilization. Taken together, the results suggest that the mechanism of ATP utilization by ClpB is adaptable and can vary depending on the specific substrate and the presence of the DnaK system.
Content may be subject to copyright.
Coupling ATP utilization to protein remodeling
by ClpB, a hexameric AAAprotein
Joel R. Hoskins
1
, Shannon M. Doyle
1
, and Sue Wickner
2
Laboratory of Molecular Biology, National Cancer Institute, National Institutes of Health, Bethesda, MD 20892
Contributed by Sue Wickner, October 22, 2009 (sent for review August 14, 2009)
ClpB and Hsp104 are members of the AAA(ATPases associated
with various cellular activities) family of proteins and are molecular
machines involved in thermotolerance. They are hexameric pro-
teins containing 12 ATP binding sites with two sites per protomer.
ClpB and Hsp104 possess some innate protein remodeling activi-
ties; however, they require the collaboration of the DnaK/Hsp70
chaperone system to disaggregate and reactivate insoluble aggre-
gated proteins. We investigated the mechanism by which ClpB
couples ATP utilization to protein remodeling with and without the
DnaK system. When wild-type ClpB, which is unable to remodel
proteins alone in the presence of ATP, was mixed with a ClpB
mutant that is unable to hydrolyze ATP, the heterohexamers
surprisingly gained protein remodeling activity. Optimal protein
remodeling by the heterohexamers in the absence of the DnaK
system required approximately three active and three inactive
protomers. In addition, the location of the active and inactive ATP
binding sites in the hexamer was not important. The results
suggest that in the absence of the DnaK system, ClpB acts by a
probabilistic mechanism. However, when we measured protein
disaggregation by ClpB heterohexamers in conjunction with the
DnaK system, incorporation of a single inactive ClpB subunit
blocked activity, supporting a sequential mechanism of ATP utili-
zation. Taken together, the results suggest that the mechanism of
ATP utilization by ClpB is adaptable and can vary depending on the
specific substrate and the presence of the DnaK system.
DnaJ DnaK GrpE protein disaggregation
Bacterial ClpB and yeast Hsp104 are ATP-dependent protein
remodeling machines that function to disaggregate protein
aggregates and reactivate proteins after extreme stress condi-
tions (1–3). In the cell, ClpB acts in conjunction with the DnaK
chaperone system and Hsp104 acts with the Hsp70 chaperone
system (4, 5). DnaK and Hsp70 are members of another large,
ubiquitous family of ATP-dependent molecular chaperones that
mediate protein reactivation and remodeling in concert with two
cochaperones, DnaJ and GrpE in prokaryotes and Hsp40 and
NEF in eukaryotes (6). Alone, neither the DnaK/Hsp70 chap-
erone system nor ClpB/Hsp104 has the ability to reactivate large
insoluble aggregates.
ClpB/Hsp104 exists as a hexameric ring with an axial channel
(7–10). Each protomer contains two AAA(ATPases associ-
ated with various cellular activities) nucleotide-binding domains
separated by a hinge region and preceded by an N-terminal
domain (1, 7). The two AAAdomains contain characteristic
motifs, including Walker A and B and sensor-1 and -2 motifs, as
well as an arginine finger (11, 12). Situated in the first AAA
domain is a long coiled-coil region, referred to as the middle
domain, which is unique to ClpB, Hsp104, and their homologs.
In vitro ClpB/Hsp104 solubilizes and reactivates protein ag-
gregates in ATP-dependent reactions in collaboration with the
DnaK/Hsp70 chaperone system (1–3). Although the roles of the
two chaperone systems in disaggregation are not fully under-
stood, it is likely that ClpB/Hsp104 is the primary protein
disaggregating machine, and DnaK/Hsp70 facilitates the inter-
action of ClpB/Hsp104 with aggregates (1–3). However, DnaK/
Hsp70 may have additional functions.
Evidence suggests that ClpB/Hsp104 acts by forcibly extracting
polypeptides from aggregates and translocating the unfolded
regions of polypeptides through its axial channel (1–3). This
substrate translocation mechanism has been established for
ClpA and ClpX, two Clp ATPases that interact with a proteolytic
component, ClpP (13). Substrates are unfolded and threaded
through the central channels of ClpA and ClpX directly into the
chamber of ClpP where degradation occurs. Work from Bukau
and colleagues, in which ClpB and Hsp104 were engineered to
contain the ClpP interaction loop from ClpA, supports a similar
mechanism for ClpB and Hsp104 (14, 15).
The demonstration that ClpB and Hsp104 have an innate
ability to unfold and activate proteins in the absence of the
DnaK/Hsp70 chaperone system provides additional support for
an unfolding and translocation mechanism (16). However, ATP
alone is ineffective in promoting these protein remodeling
activities. Instead, mixtures of ATP and ATP
S, a slowly hy-
drolyzed ATP analog, are required for remodeling, suggesting
that with the mixture of nucleotides substrates can be held (a
function requiring ATP binding and supported by ATP
S) and
unfolded and translocated (functions requiring ATP hydrolysis).
To understand how ClpB couples ATP binding and hydrolysis
to protein remodeling, we investigated the contribution of the 12
ATP binding sites to the overall chaperone activity of ClpB both
alone and in the presence of the DnaK chaperone system.
Results
ClpB Hexamers with a Balance of Active and Inactive Nucleotide
Binding Sites Are Required for Optimal Protein Remodeling in the
Absence of the DnaK System. The 12 ATP binding sites of hex-
americ ClpB are arranged in two rings of six, with each protomer
providing one nucleotide-binding site to each ring, referred to
here as Ring-1 and Ring-2 (Fig. 1A). Ring-1 of the hexamer is
comprised of the N-terminal ATP binding sites, and Ring-2 is
comprised of the C-terminal ATP-binding sites.
Previous observations demonstrated that ATP hydrolysis at
12 ATP binding sites of the ClpB hexamer is required to elicit
the innate protein remodeling activity of ClpB, and AT P hy-
drolysis at all 12 sites prohibits activity (16). In the work
presented here, we used a green fluorescent protein (GFP)
fusion protein containing a C-terminal 15 aa peptide, GFP-15,
as a substrate for remodeling. We measured protein unfolding by
monitoring the decrease in fluorescence with time in the pres-
ence of a mutant GroEL (17) that binds unfolded proteins but
does not release them. ClpB wild-type, ClpB
(wt)
, was unable to
unfold GFP-15 in the presence of ATP (Fig. 1B). However, when
mixtures of ATP and ATP
S were used in a 1:1 ratio, there was
a large decrease in GFP fluorescence (Fig. 1C). Thus, if some
Author contributions: J.R.H., S.M.D., and S.W. designed research; J.R.H. and S.M.D. per-
formed research; J.R.H. and S.M.D. contributed new reagents/analytic tools; J.R.H., S.M.D.,
and S.W. analyzed data; and J.R.H., S.M.D., and S.W. wrote the paper.
The authors declare no conflict of interest.
1J.R.H. and S.M.D. contributed equally to this work.
2To whom correspondence should be addressed. E-mail: wickners@mail.nih.gov.
This article contains supporting information online at www.pnas.org/cgi/content/full/
0911937106/DCSupplemental.
www.pnas.orgcgidoi10.1073pnas.0911937106 PNAS
December 29, 2009
vol. 106
no. 52
22233–22238
BIOCHEMISTRY
ATP binding sites hydrolyze ATP slowly, ClpB
(wt)
activity is
elicited.
Single Walker B mutants (Fig. 1 A), with either an E279A
substitution in Ring-1, ClpB
(B1)
, or an E678A substitution in
Ring-2, ClpB
(B2)
, unfolded GFP-15 with ATP alone and were
inhibited by ATP
S, supporting the conclusion that hydrolysis by
12 ATP binding sites elicits activity (Fig. 1 Band C). These
mutants contain six wild-type nucleotide-binding sites per hex-
amer and six mutant sites that are able to bind but not hydrolyze
ATP (18). A double Walker B mutant (Fig. 1 A) with both E279A
and E678A substitutions, ClpB
(B1,B2)
, was unable to catalyze
unfolding of GFP-15 with ATP or mixtures of ATP and ATP
S,
since all of its nucleotide-binding sites are defective in ATP
hydrolysis (19) (Fig. 1 Band C). Therefore, ClpB can perform
protein unfolding when there is only one active nucleotide-
binding site per protomer, and the active nucleotide-binding sites
are all in Ring-1 or Ring-2 of the hexamer. These results also
show that hydrolysis by the two rings does not need to be coupled
for protein remodeling by ClpB alone.
To investigate whether the intrinsic remodeling activity of
ClpB demands that all six of the active sites reside in Ring-1 or
Ring-2 of the hexamer, or if the active sites can be distributed
between the two rings, we performed subunit mixing experi-
ments. Recent studies have shown that the ClpB hexamer is a
dynamic complex with subunits reshuffling on a time scale of a
minute (20, 21). Thus, in subunit mixing experiments, hetero-
hexamers are expected to be represented by a binomial distri-
bution that varies as a function of the molar ratio of each subunit
present, assuming subunits have an equal ability to be incorpo-
rated into hexamers (Fig. 1D). We mixed the double Walker B
mutant, ClpB
(B1,B2)
, with ClpB
(wt)
(shown schematically in Fig.
1E) and measured GFP-15 unfolding in the presence of ATP as
the sole nucleotide. Surprisingly, we saw that the heterohexam-
ers were active, although each protein alone was inactive in
GFP-15 unfolding (Fig. 1F). By var ying the ratio of ClpB
(wt)
and
ClpB
(B1,B2)
while holding the total ClpB concentration constant,
we observed maximal activity when there were approx imately
three active sites in each ring of the hexamer [50% ClpB
(wt)
and
50% ClpB
(B1,B2)
] (Fig. 1 Fand G). Therefore, ClpB is able to
catalyze protein remodeling when there are both active and
inactive sites in Ring-1 and Ring-2 of the hexamer and the
subunits contain either two active or two inactive nucleotide-
binding sites.
We extended these observations with ClpB
(B1,B2)
and ClpB
(wt)
heterohexamers by testing two other protein remodeling reac-
tions. In a disaggregation assay, mixtures of ClpB
(B1,B2)
and
ClpB
(wt)
were able to reactivate heat-inactivated GFP in the
presence of ATP alone (Fig. 2A). Maximal reactivation was
ClpB(wt)
(side view of hexamer) ClpB(B1)
Ring-1
Ring-2
ClpB(B2)
A
BC
Sites active for ATP
hydrolysis (red)
2 nucleotide-binding
sites in a protomer (connection
shown by white oval)
Sites able to bind but not
hydrolyze ATP (blue)
1:1 mixture of
ClpB(wt) and ClpB(B1,B2)
ClpB(B1)
ClpB(B2)
ClpB(wt)
ClpB(B1,B2)
ClpB(B1,B2)
ClpB(wt):ClpB(B1,B2)
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
ClpB(B1)
ClpB(B2)
ClpB(wt)
ClpB(B1,B2)
0 10 20 30 40 0 10 20 30 40
Time (min) Time (min)
E
Time (min)
0 10 20 30 40 50 60
Fluorescence intensity (AU)
Fluorescence intensity (AU)
Fluorescence intensity (AU)
ATP and ATP S
ATP
FATP
0:6
1:5
6:0
5:1
4:2
3:3
2:4
G
Mutant (%)
0 20 40 60 80 100
Population (fraction)
1.0
0.8
0.6
0.4
0.2
0.0
D
GFP-15 unfolded (min-1)
0 20 40 60 80 1 00
0.03
0.02
0.01
0.00
ClpB(wt) (%)
ClpB(B1,B2) (%)
100 80 60 40 20 0
Fig. 1. Optimal protein unfolding by ClpB requires that ClpB hexamers have both active and inactive ATP hydrolytic sites. (A) Diagram illustrating the location
of the 12 ATP binding sites in hexamers of ClpB(wt), single Walker B mutants in the first, ClpB(B1), or second, ClpB(B2), nucleotide-binding domain and a double
Walker B mutant, ClpB(B1,B2). The 12 sites are superimposed on a model of a ClpB hexamer generated from the crystal structure of a ClpB monomer (39). (Band
C) Protein unfolding of GFP-15 by ClpB(wt), ClpB(B1), ClpB(B2), and ClpB(B1,B2) in the presence of ATP (B) or ATP and ATP
S in a 1:1 ratio (C) as described in the Methods.
(D) Mathematical model, generated as described in ref. 20, depicting the theoretical populations of wild-type and mutant ClpB heterohexamers that contain
no (black), one (aqua), two (red), three (green), four (purple), five (orange), or six (blue) mutant subunits as a function of percent mutant. (E) Diagram showing
representative locations of wild-type (red) and Walker B mutant ATP binding sites (blue) for a heterohexamer of ClpB(wt) and ClpB(B1,B2) in a 1:1 ratio. (F) GFP-15
unfolding in the presence of ATP using mixtures of ClpB(wt) and ClpB(B1,B2) in various ratios. (G) The rate of GFP-15 unfolding by mixtures of ClpB(wt) and ClpB(B1,B2)
plotted as a function of percent ClpB(B1,B2) in the mixture. In B,C, and F, representative experiments of three replicates are shown. In G, data are means SEM
(n3). Some error bars are covered by the plot symbols.
22234
www.pnas.orgcgidoi10.1073pnas.0911937106 Hoskins et al.
observed with a 1:1 ratio of the two ClpB proteins. Mixtures of
ClpB
(B1,B2)
and ClpB
(wt)
were also tested in the RepA activation
assay in which inactive RepA dimers are converted to monomers
that bind DNA with high affinity (22). In this assay as well,
maximal activation was seen with a 1:1 mixture of the two
proteins (Fig. 2A). Thus, in additional protein remodeling
reactions, heterohexamers of ClpB
(B1,B2)
and ClpB
(wt)
are func-
tional, while neither protein separately exhibits significant
activity.
When ATP hydrolysis was measured, approximately 2-fold
higher steady-state hydrolysis was observed with a 1:1 mixture of
ClpB
(wt)
and ClpB
(B1,B2)
compared to ClpB
(wt)
alone (Fig. 2B)
(19). This suggests that the activity of wild-type subunits within
the heterohexamer is stimulated. For example, at a 1:1 ratio of
ClpB
(wt)
to ClpB
(B1,B2)
, the ATPase activity of the wild-type
subunits must increase approximately 4-fold to account for the
observed increase in ATP hydrolysis. For comparison, when we
measured ATP hydrolysis by mixtures of ClpB
(wt)
and ClpB
(B1,B2)
in the presence of ATP and ATP
S, hydrolysis decreased
linearly as the percent of ClpB
(B1,B2)
in the mixture was increased
(Fig. 2B). AT P hydrolysis by ClpB
(wt)
alone was approximately
2-fold higher with a 1:1 mixture of ATP and ATP
S than with
ATP alone (Fig. 2 B) (16). Thus, conditions optimal for protein
remodeling are also optimal for ATP hydrolysis.
Our results with the single Walker B mutants and heterohex-
amers of ClpB
(wt)
and ClpB
(B1,B2)
suggest that for remodeling
activity by ClpB alone: (i) protomers need not contain a pair of
active sites as long as there are a total of six active sites in either
Ring-1 or Ring-2 of the hexamer, and (ii) there need not be six
functional sites in a single ring as long as there are approximately
six active sites in the hexamer. Together these results indicate
that the innate remodeling activity of ClpB might simply require
a total of approximately six active AT P hydrolytic sites, inde-
pendent of the location of the sites within the subunit or within
the hexamer.
To test this possibility, we measured protein unfolding by
heterohexamers of the two single Walker B mutants, ClpB
(B1)
and ClpB
(B2)
(Fig. 3A). When ClpB
(B1)
and ClpB
(B2)
were mixed
in various ratios while keeping the total concentration of ClpB
constant, there was no significant gain or loss of function (Fig.
3B). Interpretation of these results depends upon ClpB
(B1)
and
ClpB
(B2)
forming heterohexamers, which is likely, since each
single Walker B mutant is capable of forming heterohexamers
with both ClpB
(wt)
and ClpB
(B1,B2)
(shown in Fig. 4 Band Dand
Fig. 5B). The results suggest that ClpB is active in protein
remodeling when the hexamer contains six active AT P hydrolytic
sites irrespective of the position of the active sites within the
hexamer.
Mechanism of ATP Utilization by ClpB in the Absence of the DnaK
System. Our observations that heterohexamers of ClpB
(wt)
and
ClpB
(B1,B2)
perform remodeling activities demonstrate that pro-
tein unfolding and remodeling can be carried out with approx-
imately three active protomers per hexamer (Figs. 1 and 2).
These results rule out a concerted mechanism of ATP utilization
within a ring, where all six active sites of a ring simultaneously
bind ATP, hydrolyze ATP, and release ADP. They also rule out
a strictly sequential mechanism, where activity depends upon an
endless cycle of ATP utilization around the ring. Instead, the
observations point to either a probabilistic mechanism, involving
a random order of mutually independent hydrolysis events or a
semisequential mechanism, in which the sequence and timing of
ATP hydrolysis operates in a sequential manner that proceeds
around the ring with hydrolysis pausing occasionally and then
resuming at another site in the ring.
To further explore the mechanism of ATP utilization by ClpB,
we varied the number of active sites in one ring while maintaining
six inactive sites in the other ring. We first decreased the number
of active sites in Ring-2 by increasing the percentage of
ClpB
(B1,B2)
in mixtures with ClpB
(B1)
(Fig. 4A). We observed that
incorporation of one or two inactive sites in Ring-2 had little
effect on hexamer activity, while incorporation of four or five
inactive sites significantly inhibited activity (Fig. 4B). We next
changed the number of active sites in Ring-1 by varying the ratio
of ClpB
(B1,B2)
in mixtures with ClpB
(B2)
(Fig. 4C) and observed
similar results (Fig. 4D). Taken together, these experiments show
that ClpB is able to function in protein remodeling with one
ATP hydrolyzed (nmol)
100 80 60 40 20 0
0 20 40 60 80 100
B
A
GFP reactivation (% min-1)
0.06
0.04
0.02
0.00
3
2
1
0
RepA activation
(fmol oriP1 DNA bound)
100 80 60 40 20 0
0 20 40 6 0 80 100
ClpB(wt)
(%)
ClpB(B1,B2) (%)
ClpB(wt)
(%)
ClpB(B1,B2) (%)
25
20
15
10
5
0
ATP
ATP + ATP S
GFP
reactivation
RepA
activation
Fig. 2. ClpB hexamers containing approximately six active ATP hydrolytic
sites are required for optimal chaperone activity. (A) Reactivation of heat-
aggregated GFP and activation of RepA was measured as described in the
Methods using various ratios of ClpB(wt) to ClpB(B1,B2). GFP reactivation rates
(left axis) and DNA binding by RepA (right axis) were plotted as a function of
percent ClpB(B1,B2).(B) ATPase activity by mixtures of ClpB(wt) and ClpB(B1,B2) in
the presence of ATP or ATP and ATP
S in a 1:1 ratio was measured as described
in the Methods. ATPase activity is plotted as a function of percent ClpB(B1,B2).
Data are means SEM (n3). Some error bars are covered by plot symbols.
A
1:1 mixture of ClpB(B1)
and ClpB(B2)
ClpB(B2) (%)
ClpB(B1) (%)
100 80 60 40 20 0
0 20 40 60 80 100
GFP-15 unfolded (relative)
Sites able to
bind but not
hydrolyze ATP
Active
sites
B
1.0
0.8
0.6
0.4
0.2
0.0
Fig. 3. Protein unfolding by mixtures of ClpB(B1) and ClpB(B2).(A) Diagram
showing a heterohexamer containing ClpB(B1) and ClpB(B2) in a 1:1 ratio. (B)
GFP-15 unfolding by mixtures of ClpB(B1) and ClpB(B2) was measured as de-
scribed in the Methods. Rates are expressed as a fraction relative to ClpB(B2)
alone (0.060 0.002 min1) and plotted as a function of percent ClpB(B1) in the
mixture. Data are means SEM (n3). Some error bars are covered by plot
symbols.
Hoskins et al. PNAS
December 29, 2009
vol. 106
no. 52
22235
BIOCHEMISTRY
inactive ring and approximately three or more active sites in the
other ring of the hexamer (Fig. 4 Band D). Thus, the results best
support a probabilistic or a semisequential mechanism of ATP
utilization by the ClpB rings. However, six active sites per
hexamer are optimal for protein remodeling in the absence of the
DnaK system (Figs. 1 Fand Gand 2–4).
Mechanism of ATP Utilization by ClpB When Disaggregating Sub-
strates That Require the DnaK System. To test whether the mech-
anism of ATP utilization by ClpB is the same when ClpB works
on aggregates with the DnaK system as when it works alone, we
monitored disaggregation by mixtures of ClpB
(wt)
and
ClpB
(B1,B2)
in conjunction with DnaK, DnaJ, and GrpE. We used
aggregated malate dehydrogenase (MDH) as a substrate, since
previous studies showed that MDH disaggregation by ClpB in
combination with the DnaK system was approximately 20-fold
greater than disaggregation by either chaperone alone (23, 24).
As the percent of ClpB
(B1,B2)
in mixtures with ClpB
(wt)
was
increased, we observed a rapid exponential decrease in MDH
disaggregation, compared to the linear decrease expected if a
probabilistic mechanism was used (Fig. 5A). The loss of activity
when the percentage of ClpB
(B1,B2)
is low indicates that incor-
poration of approximately one protomer with two hydrolytically
defective ATP binding sites blocks disaggregation of MDH. Our
results indicate a striking departure from the protein remodeling
and ATP-ase activities seen in the absence of the DnaK system
(Figs. 1 Fand Gand 2 Aand B) and from the ATPase data (Fig.
2B). They suggest that for disaggregation of protein aggregates
in collaboration with the DnaK system, cooperative interactions
between all six ClpB subunits are required. Similar inter-subunit
coupling during remodeling has been observed for Thermus
thermophilus ClpB using another substrate, aggregated
-gluco-
sidase (20).
We next measured the effects of single nucleotide binding site
mutants in mixtures with ClpB
(wt)
. We used the two single
Walker B mutants discussed above, ClpB
(B1)
and ClpB
(B2)
.In
addition, we tested a ClpB Walker A mutant, ClpB
(A2)
, with a
K611T substitution in Ring-2 that renders it unable to bind ATP
in Ring-2.* All three mutants are defective in disaggregation of
MDH in combination with the DnaK system (23, 25) (Fig. 5 B
and C). As the percentage of ClpB
(B1)
, ClpB
(B2)
, or ClpB
(A2)
in
mixtures with ClpB
(wt)
was increased, disaggregation of MDH
decreased more rapidly than the linear decrease expected for a
probabilistic mechanism of action (Fig. 5 Band C). However,
incorporating approximately two inactive sites in the same ring
inhibited disaggregation less than incorporating approximately
one protomer with two inactive sites (Fig. 5 Aand B). These
observations suggest that for ClpB to work on substrates that
require collaboration with the DnaK system, both inter- and
intra-ring communication between ClpB subunits is important.
Together the data support a sequential or semisequential mech-
anism of ATP utilization within each ring and within the
hexamer. Moreover, they imply that the mechanism of ATP
utilization by ClpB can vary depending on the substrate and the
requirement for the DnaK system.
To address the question of whether the specific aggregate
influences ATP utilization requirements, we compared disag-
gregation of aggregated MDH to that of heat-aggregated GFP-
38, a GFP fusion protein containing a C-terminal 38 aa peptide.
Aggregated GFP-38, like MDH, required the combination of
ClpB
(wt)
and the DnaK system for reactivation, as measured by
the regain of GFP fluorescence with time (Fig. 6A). ClpB
(B1)
and
ClpB
(B2)
were each able to reactivate GFP-38 at a low rate in the
presence of the DnaK system, although neither mutant was able
to detectably disaggregate MDH (Figs. 5Band 6 A). These results
suggest that with some aggregates, six active ATP binding sites
are sufficient to carry out limited disaggregation in conjunction
with the DnaK system.
To extend the comparison of the two aggregates, we measured
disaggregation of GFP-38 by mixtures of ClpB
(B1,B2)
and Clp-
B
(wt)
. Disaggregation activity decreased exponentially as the
ratio of ClpB
(B1,B2)
to ClpB
(wt)
in the mixture increased (Fig. 6B).
The data provide further support that cooperative interactions
between subunits are important and that ATP utilization by
ClpB is likely through a sequential or semisequential mechanism
when ClpB acts in protein disaggregation in conjunction with the
DnaK system.
In summary the results presented here suggest that the
mechanism of ATP utilization by the two rings of ClpB can
change dependent upon the specific substrate and the require-
ment for the DnaK chaperone system.
Discussion
The surprising result that the heterohexamer of ClpB
(wt)
and
ClpB
(B1,B2)
is active under conditions where neither homohex-
amer is active shows that modulation of the hydrolytic cycle of
ClpB elicits protein remodeling. Moreover, approximately six
active nucleotide-binding sites are required for maximal pro-
tein remodeling activity by ClpB alone (Figs. 1 and 2). These
observations are consistent with our previous findings that
ClpB and Hsp104 perform protein remodeling alone in the
presence of mixtures of ATP and ATP
S (16). Together the
data suggest that remodeling activity is elicited when some
sites hydrolyze ATP and other sites bind ATP, but hydrolyze
*Unlike a Walker A mutant with a K611T substitution in Ring-2, a Walker A mutant with a
substitution in Ring-1, K212T, is unable to form hexamers (22, 27). For this reason,
ClpB(K212T) and a double Walker A mutant were not used.
Fig. 4. Protein unfolding by mixtures of ClpB(B1,B2) and either ClpB(B1) or
ClpB(B2).(Aand C) Diagrams showing heterohexamers of ClpB(B1,B2) with
ClpB(B1) (A) or ClpB(B2) (C) in a 1:1 ratio. (Band D) GFP-15 unfolding by mixtures
of ClpB(B1,B2) and ClpB(B1) (B) or ClpB(B2) (D) as described in the Methods.
Unfolding rates are expressed as a fraction relative to ClpB(B1) (0.035 0.001
min1)(B) or ClpB(B2) (0.060 0.002 min1)(D) alone and plotted as a function
of percent ClpB(B1,B2) in the mixture. Data are means SEM (n3). Some error
bars are hidden by plot symbols.
22236
www.pnas.orgcgidoi10.1073pnas.0911937106 Hoskins et al.
it slowly or not at all. One interpretation is that the sites that
bind but do not hydrolyze ATP help stabilize interactions
between ClpB and the substrate as well as interactions among
ClpB protomers in the hexamer, while the sites that hydrolyze
ATP are necessary for substrate unfolding and translocation.
The balance between the two types of sites influences the
efficiency of protein remodeling. However, modulating the
ATP hydrolytic cycle is not sufficient to elicit the full activity
of ClpB, since in vivo, mutants defective in AT P hydrolysis in
either ring are also defective for thermotolerance (18). One
possibility is that the substrate and the DnaK system regulate
ATP utilization, consistent with our previous results showing
that ATP hydrolysis is stimulated by the combined presence of
ClpB, the DnaK system, and aggregated substrate (23).
Interestingly, our results suggest that the number of hydro-
lytically active sites sufficient to perform remodeling activity
and the mechanism of ATP utilization can vary depending on
the substrate and the involvement of the DnaK system. When
ClpB acts in the absence of the DnaK system, the mechanism
of ATP utilization by the two rings is likely probabilistic,
although a semisequential mechanism is also possible. This is
based on experiments showing that approximately three active
protomers per hexamer are optimal for remodeling activity. In
addition, approximately three active ATP hydrolytic sites in a
ring are sufficient for protein remodeling when the other ring
contains six sites defective in ATP hydrolysis. In contrast,
when ClpB acts on aggregated substrates that require the
DnaK system, ClpB protomers must work together in a
cooperative fashion, supporting a sequential mechanism of
ATP utilization. However, the observations that heterohex-
amers can tolerate protomers with one active and one inactive
ATP binding site provide evidence for a semisequential mech-
anism of ATP utilization within the separate rings.
Probabilistic, semisequential, and sequential modes of ac-
tion have been proposed for other AA Aproteins (10, 26, 27).
For example, another hexameric Clp protein, ClpX, has been
shown to function in a probabilistic manner by Martin, Baker,
and Sauer (26). In contrast, Saibil, Lindquist, and colleagues
recently proposed that Hsp104, the yeast homolog of ClpB,
uses a sequential mechanism, based on electron microscopic
data showing asymmetry in the hexameric model (10). Crystal
structures showing asymmetric conformations have been ob-
served for several AAAhelicases, suggesting that those
proteins act by a sequential mechanism (27). Additionally,
subunit mixing studies provided evidence for a semisequential
mechanism of action by MCM and RuvB and a sequential
mechanism for T7gp4 (28–32).
While further biochemical and structural analyses of ClpB and
its interaction with the DnaK system are needed, the mechanistic
insights revealed by this study of ClpB may extend to Hsp104 and
other ClpB homologs.
Methods
Proteins and DNA. P1 RepA (22), GFP (33), GFP-15 (34), ClpB and ClpB mutants
(35), GroELtrap (17), DnaK (36), DnaJ (36), GrpE (36), and [3H]oriP1 DNA (22)
(4,475 cpm/fmol) were prepared as described. pET-GFP-38 was created by
inserting GAT and GAC codons before the multicloning site stop codon of
pET-GFP-X30-H6(37) by QuikChange (Stratagene) mutagenesis. GFP-38 was
purified as described for GFP-X30-H6(37). MDH (Roche) was labeled with 3Has
ABC
ClpB(wt) + ClpB(B1)
ClpB(wt) + ClpB(B2)
ClpB(wt) (%)
ClpB(B1,B2) (%)
0 20 40 60 80 100
ClpB(B1) or ClpB(B2) (%)
0 20 40 60 80 100
ClpB(A2) (%)
0 20 40 60 80 100
Soluble MDH (fraction)
Soluble MDH (fraction)
Soluble MDH (fraction)
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
ClpB(wt) (%)
ClpB(wt) (%)
100 80 60 40 20 0 100 80 60 40 20 0
100 80 60 40 20 0
ClpB(wt) + ClpB(B1,B2)
ClpB(wt) + ClpB(A2)
Fig. 5. Disaggregation of MDH by mixtures of ClpB(wt) and ClpB mutants in conjunction with DnaK, DnaJ, and GrpE. (AC) Recovery of soluble MDH from
heat-induced aggregates was measured as described in the Methods for mixtures of ClpB(wt) with ClpB(B1,B2) (A), ClpB(B1) (B), ClpB(B2) (B), or ClpB(A2) (C)inthe
presence of DnaK/DnaJ/GrpE. In AC, the dashed gray line represents the linear decrease expected for a probabilistic mechanism where the activity of the hexamer
is proportional to the number of wild-type subunits. The fraction of soluble MDH recovered in reactions containing 100% ClpB(wt) was set equal to one. With
conditions used, ClpB(wt) solubilized 53% 1% of the MDH. Data are means SEM (n3). Some error bars are covered by plot symbols.
No chaperone
K/J/E
ClpB(wt)
ClpB(wt) + K/J/E
ClpB(B1)
ClpB(B1) + K/J/E
ClpB(B2)
ClpB(B2) + K/J/E
A
B
GFP-38 reactivation (% min-1)
ClpB(B1,B2) (%)
GFP-38 reactivation
(% min-1)
0 1.0 2.0 3.0 4.0
3.0
2.0
1.0
0
0 20 4 0 60 80 100
ClpB(wt) (%)
100 80 60 40 20 0
ClpB(wt) + ClpB(B1,B2)
Fig. 6. Disaggregation of GFP-38 by mixtures of ClpB(wt) and ClpB mutants in
conjunction with DnaK, DnaJ, and GrpE. (A) The rate of reactivation of
heat-inactivated GFP-38 by ClpB(wt) or ClpB Walker B mutants was measured
with and without DnaK/DnaJ/GrpE as described in the Methods.(B) The rate
of GFP-38 reactivation by mixtures of ClpB(wt) and ClpB(B1,B2) with DnaK/DnaJ/
GrpE was measured as described in the Methods and plotted as a function of
percent ClpB(B1,B2). The dashed gray line represents the linear decrease ex-
pected for a probabilistic mechanism where the activity of the hexamer is
proportional to the number of wild-type subunits. Data are means SEM (n
3). Some error bars are hidden by plot symbols.
Hoskins et al. PNAS
December 29, 2009
vol. 106
no. 52
22237
BIOCHEMISTRY
described in ref. 38. Protein concentrations given are for monomeric GFP,
GFP-15, GFP-38, MDH, and DnaK; dimeric RepA, DnaJ, and GrpE; hexameric
ClpB; and tetradecameric GroELtrap.
Assays. GFP-fusion protein unfolding (16), GFP reactivation (16), MDH disaggre-
gation (23), RepA activation (16), and ATPase (16) assays were performed as
described with slight modifications described in the SI Methods. GFP-38 reacti-
vation was performed as described in the SI Methods. For subunit mixing exper-
iments, control experiments were conducted to demonstrate that the activity of
ClpB(wt), ClpB(B1), and ClpB(B2) decreased linearly upon dilution over the range of
concentrations used.
ACKNOWLEDGMENTS. We thank Danielle Johnston, Jodi Camberg, Marika
Miot, and Olivier Genest for critical reading of the manuscript and helpful
discussions. We thank Michal Zolkiewski (Kansas State University) for the
plasmid pET20b-K611T and BK Lee (National Cancer Institute) for generating
mathematical models. This work was supported by the Intramural Research
Program of the National Institutes of Health (NIH) National Cancer Institute
Center for Cancer Research.
1. Doyle SM, Wickner S (2009) Hsp104 and ClpB: Protein disaggregating machines. Trends
Biochem Sci 34:40– 48.
2. Mogk A, Haslberger T, Tessarz P, Bukau B (2008) Common and specific mechanisms of
AAAproteins involved in protein quality control. Biochem Soc Trans 36:120–125.
3. Zolkiewski M (2006) A camel passes through the eye of a needle: Protein unfolding
activity of Clp ATPases. Mol Microbiol 61:1094–1100.
4. Parsell DA, Kowal AS, Singer MA, Lindquist S (1994) Protein disaggregation mediated
by heat-shock protein Hsp104. Nature 372:475–478.
5. Mogk A, et al. (1999) Identification of thermolabile Escherichia coli proteins: Preven-
tion and reversion of aggregation by DnaK and ClpB. EMBO J 18:6934– 6949.
6. Bukau B, Horwich AL (1998) The Hsp70 and Hsp60 chaperone machines. Cell 92:351–
366.
7. Lee S, et al. (2003) The structure of ClpB: A molecular chaperone that rescues proteins
from an aggregated state. Cell 115:229–240.
8. Lee S, Choi JM, Tsai FT (2007) Visualizing the ATPase cycle in a protein disaggregating
machine: Structural basis for substrate binding by ClpB. Mol Cell 25:261–271.
9. Wendler P, et al. (2007) Atypical AAAsubunit packing creates an expanded cavity for
disaggregation by the protein-remodeling factor Hsp104. Cell 131:1366–1377.
10. Wendler P, et al. (2009) Motor mechanism for protein threading through Hsp104. Mol
Cell 34:81–92.
11. Hanson PI, Whiteheart SW (2005) AAAproteins: Have engine, will work. Nat Rev Mol
Cell Biol 6:519–529.
12. Ogura T, Whiteheart SW, Wilkinson AJ (2004) Conserved arginine residues implicated
in ATP hydrolysis, nucleotide-sensing, and inter-subunit interactions in AAA and AAA
ATPases. J Struct Biol 146:106–112.
13. Sauer RT, et al. (2004) Sculpting the proteome with AAA() proteases and disassembly
machines. Cell 119:9–18.
14. Tessarz P, Mogk A, Bukau B (2008) Substrate threading through the central pore of the
Hsp104 chaperone as a common mechanism for protein disaggregation and prion
propagation. Mol Microbiol 68:87–97.
15. Weibezahn J, et al. (2004) Thermotolerance requires refolding of aggregated proteins
by substrate translocation through the central pore of ClpB. Cell 119:653–665.
16. Doyle SM, et al. (2007) Asymmetric deceleration of ClpB or Hsp104 ATPase activity
unleashes protein-remodeling activity. Nat Struct Mol Biol 14:114–122.
17. Weber-Ban EU, Reid BG, Miranker AD, Horwich AL (1999) Global unfolding of a
substrate protein by the Hsp100 chaperone ClpA. Nature 401:90–93.
18. Mogk A, et al. (2003) Roles of individual domains and conserved motifs of the AAA
chaperone ClpB in oligomerization, ATP hydrolysis, and chaperone activity. J Biol Chem
278:17615–17624.
19. Weibezahn J, Schlieker C, Bukau B, Mogk A (2003) Characterization of a trap mutant
of the AAAchaperone ClpB. J Biol Chem 278:32608–32617.
20. Werbeck ND, Schlee S, Reinstein J (2008) Coupling and dynamics of subunits in the
hexameric AAAchaperone ClpB. J Mol Biol 378:178–190.
21. Haslberger T, et al. (2008) Protein disaggregation by the AAAchaperone ClpB
involves partial threading of looped polypeptide segments. Nat Struct Mol Biol
15:641–650.
22. Wickner S, Hoskins J, McKenney K (1991) Function of DnaJ and DnaK as chaperones in
origin-specific DNA binding by RepA. Nature 350:165–167.
23. Doyle SM, Hoskins JR, Wickner S (2007) Collaboration between the ClpB AAA
remodeling protein and the DnaK chaperone system. Proc Natl Acad Sci USA
104:11138–11144.
24. Schlieker C, Tews I, Bukau B, Mogk A (2004) Solubilization of aggregated proteins by
ClpB/DnaK relies on the continuous extraction of unfolded polypeptides. FEBS Lett
578:351–356.
25. Mogk A, et al. (2003) Refolding of substrates bound to small Hsps relies on a disag-
gregation reaction mediated most efficiently by ClpB/DnaK. J Biol Chem 278:31033–
31042.
26. Martin A, Baker TA, Sauer RT (2005) Rebuilt AAAmotors reveal operating principles
for ATP-fuelled machines. Nature 437:1115–1120.
27. Enemark EJ, Joshua-Tor L (2008) On helicases and other motor proteins. Curr Opin
Struct Biol 18:243–257.
28. Mezard C, Davies AA, Stasiak A, West SC (1997) Biochemical properties of RuvBD113N:
A mutation in helicase motif II of the RuvB hexamer affects DNA binding and ATPase
activities. J Mol Biol 271:704–717.
29. Hishida T, Iwasaki H, Han YW, Ohnishi T, Shinagawa H (2003) Uncoupling of the ATPase
activity from the branch migration activity of RuvAB protein complexes containing
both wild-type and ATPase-defective RuvB proteins. Genes Cells 8:721–730.
30. Moreau MJ, McGeoch AT, Lowe AR, Itzhaki LS, Bell SD (2007) ATPase site architecture
and helicase mechanism of an archaeal MCM. Mol Cell 28:304–314.
31. Crampton DJ, Mukherjee S, Richardson CC (2006) DNA-induced switch from indepen-
dent to sequential dTTP hydrolysis in the bacteriophage T7 DNA helicase. Mol Cell
21:165–174.
32. Singleton MR, Sawaya MR, Ellenberger T, Wigley DB (2000) Crystal structure of T7 gene
4 ring helicase indicates a mechanism for sequential hydrolysis of nucleotides. Cell
101:589– 600.
33. Hoskins JR, Kim SY, Wickner S (2000) Substrate recognition by the ClpA chaperone
component of ClpAP protease. J Biol Chem 275:35361–35367.
34. Hoskins JR, Yanagihara K, Mizuuchi K, Wickner S (2002) ClpAP and ClpXP degrade
proteins with tags located in the interior of the primary sequence. Proc Natl Acad Sci
USA 99:11037–11042.
35. Zolkiewski M, Kessel M, Ginsburg A, Maurizi MR (1999) Nucleotide-dependent oli-
gomerization of ClpB from Escherichia coli.Protein Sci 8:1899–1903.
36. Skowyra D, Wickner S (1993) The interplay of the GrpE heat shock protein and Mg2
in RepA monomerization by DnaJ and DnaK. J Biol Chem 268:25296–25301.
37. Hoskins JR, Wickner S (2006) Two peptide sequences can function cooperatively to
facilitate binding and unfolding by ClpA and degradation by ClpAP. Proc Natl Acad Sci
USA 103:909–914.
38. Hoskins JR, Pak M, Maurizi MR, Wickner S (1998) The role of the ClpA chaperone in
proteolysis by ClpAP. Proc Natl Acad Sci USA 95:12135–12140.
39. Diemand AV, Lupas AN (2006) Modeling AAAring complexes from monomeric
structures. J Struct Biol 156:230–243.
22238
www.pnas.orgcgidoi10.1073pnas.0911937106 Hoskins et al.
... Allosteric interactions of the M domain and the NBDs. The binding and hydrolysis of ATP have been shown to affect the activity of ClpB in distinct ways 28,36,[55][56][57][58][59] . Each of the two nucleotide binding sites, NBD1 and NBD2, contains conserved Walker A and Walker B motifs 60,61 . ...
... Here we found both Walker A mutants to be well assembled under our experimental conditions, which suggests that both NBDs play a role in hexamer stabilization ( Supplementary Fig. 3). However, we clearly observe that NBD2 contributes more a b c significantly to the machine activity ( Fig. 3a Fig. 13), but as expected 55,69 , no ATPase activity and therefore no disaggregation activity was observed DnaK and substrate binding regulate M domain dynamics. DnaK is the main component of the co-chaperone system in the disaggregation process, which has been shown to act both upstream and downstream of ClpB 25,30,63,70 . ...
Article
Full-text available
Large protein machines are tightly regulated through allosteric communication channels. Here we demonstrate the involvement of ultrafast conformational dynamics in allosteric regulation of ClpB, a hexameric AAA+ machine that rescues aggregated proteins. Each subunit of ClpB contains a unique coiled-coil structure, the middle domain (M domain), proposed as a control element that binds the co-chaperone DnaK. Using single-molecule FRET spectroscopy, we probe the M domain during the chaperone cycle and find it to jump on the microsecond time scale between two states, whose structures are determined. The M-domain jumps are much faster than the overall activity of ClpB, making it an effectively continuous, tunable switch. Indeed, a series of allosteric interactions are found to modulate the dynamics, including binding of nucleotides, DnaK and protein substrates. This mode of dynamic control enables fast cellular adaptation and may be a general mechanism for the regulation of cellular machineries.
... A limited, Hsp70-independent disaggregation activity of ClpB/Hsp104 has been reported in presence of ATP/ATPγS mixtures (Doyle et al., 2007a,b;Hoskins et al., 2009). The gained disaggregation activity is, however, limited and also dependent on the aggregated model protein. ...
Article
Full-text available
Bacteria as unicellular organisms are most directly exposed to changes in environmental growth conditions like temperature increase. Severe heat stress causes massive protein misfolding and aggregation resulting in loss of essential proteins. To ensure survival and rapid growth resume during recovery periods bacteria are equipped with cellular disaggregases, which solubilize and reactivate aggregated proteins. These disaggregases are members of the Hsp100/AAA+ protein family, utilizing the energy derived from ATP hydrolysis to extract misfolded proteins from aggregates via a threading activity. Here, we describe the two best characterized bacterial Hsp100/AAA+ disaggregases, ClpB and ClpG, and compare their mechanisms and regulatory modes. The widespread ClpB disaggregase requires cooperation with an Hsp70 partner chaperone, which targets ClpB to protein aggregates. Furthermore, Hsp70 activates ClpB by shifting positions of regulatory ClpB M-domains from a repressed to a derepressed state. ClpB activity remains tightly controlled during the disaggregation process and high ClpB activity states are likely restricted to initial substrate engagement. The recently identified ClpG (ClpK) disaggregase functions autonomously and its activity is primarily controlled by substrate interaction. ClpG provides enhanced heat resistance to selected bacteria including pathogens by acting as a more powerful disaggregase. This disaggregase expansion reflects an adaption of bacteria to extreme temperatures experienced during thermal based sterilization procedures applied in food industry and medicine. Genes encoding for ClpG are transmissible by horizontal transfer, allowing for rapid spreading of extreme bacterial heat resistance and posing a threat to modern food production.
... The apparent IC50 for DBeQ is an order of magnitude lower than the apparent Kd ( Figure 1, 2, 3, and Supplementary Table 1), which suggests that partial saturation of the DBeQ sites in the ClpB hexamer is sufficient for a strong inhibition. Indeed, it has been observed that incorporation of a single inactive subunit into a hexameric ClpB blocks aggregate reactivation (50). ...
Article
Full-text available
The ClpB/DnaK bi-chaperone system reactivates aggregated cellular proteins and is essential for survival of bacteria, fungi, protozoa, and plants under stress. AAA+ ATPase ClpB is a promising target for the development of antimicrobials, because a loss of its activity is detrimental for survival of many pathogens and no apparent ClpB orthologs are found in metazoans. We investigated ClpB activity in the presence of several compounds that were previously described as inhibitor leads for the human AAA+ ATPase p97, an anti-tumor target. We discovered that N2,N4-dibenzylquinazoline-2,4-diamine (DBeQ), the least potent among the tested p97 inhibitors, binds to ClpB with a Kd~60 μM and inhibits the casein-activated, but not the basal ATPase activity of ClpB with an IC50~5 μM. The remaining p97 ligands, which displayed a higher affinity towards p97, did not affect the ClpB ATPase. DBeQ also interacted with DnaK with a Kd~100 μM, did not affect the DnaK ATPase, but inhibited the DnaK chaperone activity in vitro. DBeQ inhibited the reactivation of aggregated proteins by the ClpB/DnaK bi-chaperone system in vitro with an IC50~5 μM and suppressed the growth of cultured E. coli. The DBeQ-induced loss of E. coli proliferation was exacerbated by heat shock, but was nearly eliminated in a ClpB-deficient E. coli strain, which demonstrates a significant selectivity of DBeQ towards ClpB in cells. Our results provide chemical validation of ClpB as a target for developing novel antimicrobials. We identified DBeQ as a promising lead compound for structural optimization aimed at selective targeting of ClpB and/or DnaK.
... ; https://doi.org/10.1101/189316 doi: bioRxiv preprint model is in agreement with mechanisms proposed for other AAA+ ATPases (Enemark and Joshua-Tor 2006, Augustin 2009, Hoskins 2009, Gates 2017, Monroe 2017, Ripstein 2017). ...
Preprint
Full-text available
We present the first atomic model of a substrate-bound inner mitochondrial membrane AAA+ quality control protease, YME1. Our ~3.4 Å cryo-EM structure reveals how the ATPases form a closed spiral staircase encircling an unfolded substrate, directing it toward the flat, symmetric protease ring. Importantly, the structure reveals how three coexisting nucleotide states allosterically induce distinct positioning of tyrosines in the central channel, resulting in substrate engagement and translocation to the negatively charged proteolytic chamber. This tight coordination by a network of conserved residues defines a sequential, around-the-ring ATP hydrolysis cycle that results in step-wise substrate translocation. Furthermore, we identify a hinge-like linker that accommodates the large-scale nucleotide-driven motions of the ATPase spiral independently of the contiguous planar proteolytic base. These results define the first molecular mechanism for a mitochondrial inner membrane AAA+ protease and reveal a translocation mechanism likely conserved for other AAA+ ATPases.
... A mixture of ATP with the slowly hydrolyzable analog ATPgS can increase Hsp104 activities, including substrate unfolding, disaggregation, and reactivation (40). The bacterial homolog ClpB was shown to have an increased unfolding capability when the WT protein was ''doped'' with hydrolysis-deficient mutants (68). Taken together, these findings suggest that modulating ATP hydrolysis in such a way as to extend the amount of time during which Hsp104 is engaged with the peptide substrate may be the key in regulating Hsp104 activity. ...
Article
Heat shock protein (Hsp) 104 is a hexameric ATPases associated with diverse cellular activities motor protein that enables cells to survive extreme stress. Hsp104 couples the energy of ATP binding and hydrolysis to solubilize proteins trapped in aggregated structures. The mechanism by which Hsp104 disaggregates proteins is not completely understood but may require Hsp104 to partially or completely translocate polypeptides across its central channel. Here, we apply transient state, single turnover kinetics to investigate the ATP-dependent translocation of soluble polypeptides by Hsp104 and Hsp104 A503S , a potentiated variant developed to resolve misfolded conformers implicated in neurodegenerative disease. We establish that Hsp104 and Hsp104 A503S can operate as nonprocessive translocases for soluble substrates, indicating a “partial threading” model of translocation. Remarkably, Hsp104 A503S exhibits altered coupling of ATP binding to translocation and decelerated dissociation from polypeptide substrate compared to Hsp104. This altered coupling and prolonged substrate interaction likely increases entropic pulling forces, thereby enabling more effective aggregate dissolution by Hsp104 A503S .
... All recent substrate-bound structures of AAA + translocases identify a narrow channel containing a single polypeptide chain that is coordinated by a spiral array of pore-loop interactions, thus it appears unlikely that more than one strand could be accommodated in the channel (Deville et al. 2017;Gates et al. 2017;Han et al. 2017;Puchades et al. 2017). However, previous studies identified that ClpB can process internal segments of substrates (Haslberger et al. 2008) in addition to translocating from the amino-or carboxy-termini (Doyle et al. 2007;Haslberger et al. 2008;Hoskins et al. 2009). Likewise, Hsp104 can drive disaggregation via partial translocation initiated at internal segments (Haslberger et al. 2008;Sweeny et al. 2015), and can also process substrates from the amino-or carboxy-termini (Doyle et al. 2007;Sweeny et al. 2015). ...
Article
Hsp104 is a hexameric AAA+ ATPase and protein disaggregase found in yeast, which couples ATP hydrolysis to the dissolution of diverse polypeptides trapped in toxic preamyloid oligomers, phase-transitioned gels, disordered aggregates, amyloids, and prions. Hsp104 shows plasticity in disaggregating diverse substrates, but how its hexameric architecture operates as a molecular machine has remained unclear. Here, we highlight structural advances made via cryoelectron microscopy (cryo-EM) that enhance our mechanistic understanding of Hsp104 and other related AAA+ translocases. Hsp104 hexamers are dynamic and adopt open "lock-washer" spiral states and closed ring structures that envelope polypeptide substrate inside the axial channel. ATP hydrolysis-driven conformational changes at the spiral seam ratchet substrate deeper into the channel. Remarkably, this mode of polypeptide translocation is reminiscent of models for how hexameric helicases unwind DNA and RNA duplexes. Thus, Hsp104 likely adapts elements of a deeply rooted, ring-translocase mechanism to the specialized task of protein disaggregation.
... In addition, ?-casein is reported to bind E. coli ClpB with similar affinity as heat-inactivated luciferase [17]. ClpB homologs in other bacteria are known to have ATP- dependent dynamic and transient interactions with their substrates [5,34,[43][44][45] Fig. 7F) which reflects a successful interaction of both the proteins with the substrate. However, the dissociation constants for the two proteins differed appreciably, while ClpB was found to have a Kd of 0.88 ? ...
Article
Full-text available
Mycobacterium tuberculosis (M. tb) is known to persist in extremely hostile environments within host macrophages. The ability to withstand such proteotoxic stress comes from its highly conserved molecular chaperone machinery. ClpB, a unique member of the AAA+ family of chaperones, is responsible for resolving aggregates in M. tb and many other bacterial pathogens. M. tb produces two isoforms of ClpB, a full length and an N‐terminally truncated form (ClpB∆N), with the latter arising from an internal translation initiation site. It is not clear why this internal start site is conserved and what role the N‐terminal domain (NTD) of M. tb ClpB plays in its function. In the current study, we functionally characterized and compared the two isoforms of M. tb ClpB. We found the NTD to be dispensable for oligomerization, ATPase activity and prevention of aggregation activity of ClpB. Both ClpB and ClpB∆N were found to be capable of resolubilizing protein aggregates. However, the efficiency of ClpB∆N at resolubilizing higher order aggregates was significantly lower than that of ClpB. Further, ClpB∆N exhibited reduced affinity for substrates as compared to ClpB. We also demonstrated that the surface of the NTD of M. tb ClpB has a hydrophobic groove which contains four hydrophobic residues: L97, L101, F140, and V141. These residues act as initial contacts for the substrate and are crucial for stable interaction between ClpB and highly aggregated substrates.
Chapter
NMR spectroscopy has found a wide range of applications in life sciences over recent decades. Providing a comprehensive amalgamation of the scattered knowledge of how to apply high-resolution NMR techniques to biomolecular systems, this book will break down the conventional stereotypes in the use of NMR for structural studies. The major focus is on novel approaches in NMR which deal with the functional interface of either protein-protein interactions or protein-lipid interactions. Bridging the gaps between structural and functional studies, the Editors believe a thorough compilation of these studies will open an entirely new dimension of understanding of crucial functional motifs. This in turn will be helpful for future applications into drug design or better understanding of systems. The book will appeal to NMR practitioners in industry and academia who are looking for a comprehensive understanding of the possibilities of applying high-resolution NMR spectroscopic techniques in probing biomolecular interactions.
Article
Precise determination of yeast physiology is crucial in both laboratories and breweries. Our aim was to identify a sensitive indicator that could demonstrate subtle physiological differences in brewer’s yeast cultures. During 4-day ageing at 25 °C, the cell viability, intracellular adenosine triphosphate (ATP) content, and intracellular pH values changed significantly. The ATP content and intracellular pH value both correlated with viability. At 4 °C, a correlation was also observed between ATP content and viability, while no correlation was found between intracellular pH value and viability. When the relationship between intracellular ATP and viability in other strains was investigated, it was found that the ATP content could reflect both viability and vitality. When viability was significantly modified, the ATP content changed considerably more than the viability. When viability remained unchanged, the significant decrease in ATP content reflected a decrease in vitality. Taken together, it was concluded that intracellular ATP content is more sensitive than the intracellular pH value and viability; thus, this parameter can help identify subtle changes in yeast environmental responses when other commonly used indicators remain unchanged.
Article
ClpB and DnaKJE provide protection to Escherichia coli cells during extreme environmental stress. Together, this co-chaperone system can resolve protein aggregates, restoring misfolded proteins to their native form and function or solubilizing damaged proteins for removal by the cell’s proteolytic systems. DnaK is the component of the KJE system that directly interacts with ClpB. There are many hypotheses for how DnaK affects ClpB catalyzed disaggregation, each with some experimental support. Here, we build on our recent work characterizing the molecular mechanism of ClpB catalyzed polypeptide translocation by developing a stopped-flow FRET assay that allows us to detect ClpB’s movement on model polypeptide substrates in the absence or presence of DnaK. We find that DnaK induces ClpB to dissociate from the polypeptide substrate. We propose that DnaK acts as a peptide release factor, binding ClpB and causing the ClpB conformation to change to a low peptide-affinity state. Such a role for DnaK would allow ClpB to rebind to another portion of an aggregate and continue nonprocessive translocation to disrupt the aggregate.
Article
Full-text available
The protein-remodeling machine Hsp104 dissolves amorphous aggregates as well as ordered amyloid assemblies such as yeast prions. Force generation originates from a tandem AAA+ (ATPases associated with various cellular activities) cassette, but the mechanism and allostery of this action remain to be established. Our cryoelectron microscopy maps of Hsp104 hexamers reveal substantial domain movements upon ATP binding and hydrolysis in the first nucleotide-binding domain (NBD1). Fitting atomic models of Hsp104 domains to the EM density maps plus supporting biochemical measurements show how the domain movements displace sites bearing the substrate-binding tyrosine loops. This provides the structural basis for N- to C-terminal substrate threading through the central cavity, enabling a clockwise handover of substrate in the NBD1 ring and coordinated substrate binding between NBD1 and NBD2. Asymmetric reconstructions of Hsp104 in the presence of ATPgammaS or ATP support sequential rather than concerted ATP hydrolysis in the NBD1 ring.
Article
Full-text available
Genetic and biochemical studies have established that the sole function of the Escherichia coli DnaJ, DnaK, and GrpE heat shock proteins in plasmid P1 DNA replication is to convert RepA dimers to monomers. Monomers bind avidly to oriP1 DNA and initiate DNA replication. However, with purified heat shock proteins, only DnaJ, DnaK, and ATP were required for the monomerization of RepA; GrpE was not required. We have found reaction conditions that mimic the physiological situation. GrpE function is absolutely necessary for RepA activation in vitro with DnaJ and DnaK when the free Mg2+ concentration is maintained at a level of approximately 1 microM by a metal ion buffer system. EDTA or physiological metabolites, including citrate, phosphate, pyrophosphate, and ATP, all elicit the GrpE requirement. With these metal ion-buffering systems, GrpE specifically lowers the concentration of Mg2+ required for the RepA activation reaction. The absence of Mg2+ blocks activation and high levels of Mg2+ in solution bypass the requirement for GrpE but not for the other two heat shock proteins. Our results imply that GrpE facilitates the utilization of Mg2+ for an essential step in RepA activation.
Article
Full-text available
B. B. thanks members of his lab and J. Reinstein for critical reading of the manuscript and C. Gassler, T. Laufen, and S. Rudiger for figure preparation. A. H. thanks Wayne Fenton for critical reading and Zhaohui Xu for figure preparation. A. H. dedicates this work to Guenter Brueckner, always an inspiration.
Article
Molecular chaperones assist protein folding by facilitating their “forward” folding and preventing aggregation. However, once aggregates have formed, these chaperones cannot facilitate protein disaggregation. Bacterial ClpB and its eukaryotic homolog Hsp104 are essential proteins of the heat-shock response, which have the remarkable capacity to rescue stress-damaged proteins from an aggregated state. We have determined the structure of Thermus thermophilus ClpB (TClpB) using a combination of X-ray crystallography and cryo-electron microscopy (cryo-EM). Our single-particle reconstruction shows that TClpB forms a two-tiered hexameric ring. The ClpB/Hsp104-linker consists of an 85 Å long and mobile coiled coil that is located on the outside of the hexamer. Our mutagenesis and biochemical data show that both the relative position and motion of this coiled coil are critical for chaperone function. Taken together, we propose a mechanism by which an ATP-driven conformational change is coupled to a large coiled-coil motion, which is indispensable for protein disaggregation.
Article
Molecular chaperones assist protein folding by facilitating their "forward" folding and preventing aggregation. However, once aggregates have formed, these chaperones cannot facilitate protein disaggregation. Bacterial ClpB and its eukaryotic homolog Hsp104 are essential proteins of the heat-shock response, which have the remarkable capacity to rescue stress-damaged proteins from an aggregated state. We have determined the structure of Thermus thermophilus ClpB (TClpB) using a combination of X-ray crystallography and cryo-electron microscopy (cryo-EM). Our single-particle reconstruction shows that TClpB forms a two-tiered hexameric ring. The ClpB/Hsp104-linker consists of an 85 A long and mobile coiled coil that is located on the outside of the hexamer. Our mutagenesis and biochemical data show that both the relative position and motion of this coiled coil are critical for chaperone function. Taken together, we propose a mechanism by which an ATP-driven conformational change is coupled to a large coiled-coil motion, which is indispensable for protein disaggregation.
Article
Self-association of ClpB (a mixture of 95- and 80-kDa subunits) has been studied with gel filtration chromatography, analytical ultracentrifugation, and electron microscopy. Monomeric ClpB predominates at low protein concentration (0.07 mg0mL), while an oligomeric form is highly populated at >4 mg/mL. The oligomer formation is enhanced in the presence of 2 mM ATP or adenosine 5′-O-thiotriphosphate (ATPS). In contrast, 2 mMADP inhibits full oligomerization of ClpB. The apparent size of the ATP- or ATPS-induced oligomer, as determined by gel filtration, sedimentation velocity and electron microscopy image averaging, and the molecular weight, as determined by sedimentation equilibrium, are consistent with those of a ClpB hexamer. These results indicate that the oligomerization reactions of ClpB are similar to those of other Hsp100 proteins.
Article
Heat-shock protein 104 (Hsp104) and caseinolytic peptidase B (ClpB), members of the AAA+ superfamily, are molecular machines involved in disaggregating insoluble protein aggregates, a process not long ago thought to be impossible. During extreme stress they are essential for cell survival. In addition, Hsp104 regulates prion assembly and disassembly. For most of their protein remodeling activities Hsp104 and ClpB work in collaboration with the Hsp70 or DnaK chaperone systems. Together, the two chaperones catalyze protein disaggregation and reactivation by a mechanism probably involving the extraction of polypeptides from aggregates by forced unfolding and translocation through the Hsp104/ClpB central cavity. The polypeptides are then released back into the cellular milieu for spontaneous or chaperone-mediated refolding.
Article
Heat-shock proteins are normal constituents of cells whose synthesis is increased on exposure to various forms of stress. They are interesting because of their ubiquity and high conservation during evolution. Two families of heat-shock proteins, hsp60s and hsp70s, have been implicated in accelerating protein folding and oligomerization and also in maintaining proteins in an unfolded state, thus facilitating membrane transport. The Escherichia coli hsp70 analogue, DnaK, and two other heat-shock proteins, DnaJ and GrpE, are required for cell viability at high temperatures and are involved in DNA replication of phage lambda and plasmids P1 and F. These three proteins are involved in replication in vitro of P1 DNA along with many host replication proteins and the P1 RepA initiator protein. RepA exists in a stable protein complex with DnaJ containing a dimer each of RepA and DnaJ. We report here that DnaK and DnaJ mediate an alteration in the P1 initiator protein, rendering it much more active for oriP1 DNA binding.
Article
The heat-inducible members of the Hsp100 (or Clp) family of proteins share a common function in helping organisms to survive extreme stress, but the basic mechanism through which these proteins function is not understood. Hsp104 protects cells against a variety of stresses, under many physiological conditions, and its function has been evolutionarily conserved, at least from Saccharomyces cerevisiae to Arabidopsis thaliana. Homology with the Escherichia coli ClpA protein suggests that Hsp104 may provide stress tolerance by helping to rid the cell of heat-denatured proteins through proteolysis. But genetic analysis indicates that Hsp104 may function like Hsp70 as a molecular chaperone. Here we investigate the role of Hsp104 in vivo using a temperature-sensitive Vibrio harveyi luciferase-fusion protein as a test substrate. We find that Hsp104 does not protect luciferase from thermal denaturation, nor does it promote proteolysis of luciferase. Rather, Hsp104 functions in a manner not previously described for other heat-shock proteins: it mediates the resolubilization of heat-inactivated luciferase from insoluble aggregates.
Article
Many DNA helicases utilise the energy derived from nucleoside triphosphate hydrolysis to fuel their actions as molecular motors in a variety of biological processes. In association with RuvA, the E. coli RuvB protein (a hexameric ring helicase), promotes the branch migration of Holliday junctions during genetic recombination and DNA repair. To analyse the relationship between ATP-dependent DNA helicase activity and branch migration, a site-directed mutation was introduced into the helicase II motif of RuvB. Over-expression of RuvBD113N in wild-type E. coli resulted in a dominant negative UVs phenotype. The biochemical properties of RuvBD113N were examined and compared with wild-type RuvB in vitro. The single amino acid substitution resulted in major alterations to the biochemical activities of RuvB, such that RuvBD113N was defective in DNA binding and ATP hydrolysis, while retaining the ability to form hexameric rings and interact with RuvA. RuvBD113N formed heterohexamers with wild-type RuvB, and could inhibit RuvB function by affecting its ability to bind DNA. However, heterohexamers exhibited an ability to promote branch migration in vitro indicating that not all subunits of the ring need to be catalytically competent.